Sei sulla pagina 1di 14

eXPRESS Polymer Letters Vol.10, No.

11 (2016) 927940
Available online at www.expresspolymlett.com
DOI: 10.3144/expresspolymlett.2016.86

Bio-based polyurethane prepared from Kraft lignin and


modified castor oil
L. B. Tavares, C. V. Boas, G. R. Schleder, A. M. Nacas, D. S. Rosa, D. J. Santos*

Centro de Engenharia, Modelagem e Cincias Sociais Aplicadas, UFABC, Santo Andr, 09210-580 So Paulo, Brazil

Received 24 April 2016; accepted in revised form 26 June 2016

Abstract. Current challenges highlight the need for polymer research using renewable natural sources as a substitute for pe-
troleum-based polymers. The use of polyols obtained from renewable sources combined with the reuse of industrial residues
such as lignin is an important agent in this process. Different compositions of polyurethane-type materials were prepared by
combining technical Kraft lignin (TKL) with castor oil (CO) or modified castor oil (MCO1 and MCO2) to increase their re-
activity towards diphenylmethane diisocyanate (MDI). The results indicate that lignin increases the glass transition temper-
ature, the crosslinking density and improves the ultimate stress especially for those prepared from MCO2 and 30% lignin
content from 8.2 MPa (lignin free) to 23.5 MPa. Scanning electron microscopy (SEM) micrographs of rupture surface after
uniaxial tensile tests show ductile-to-brittle transition. The results show the possibility to develop polyurethane-type materials,
varying technical grade Kraft lignin content, which cover a wide range of mechanical properties (from large elastic/low
Young modulus to brittle/high Young modulus polyurethanes).

Keywords: mechanical properties, reinforcements, Kraft lignin, lignopolyurethane materials, modified castor oil

1. Introduction Castor oil is a major candidate in these replacement


Increasing concerns about depletion of petroleum- efforts due to its inherent advantages over other veg-
based resources and environmental problems caused etable oils [3]. Besides its renewability, low cost and
by petroleum-based materials have led to consider- easy availability in large quantities, castor oil is not
able efforts to develop materials based on renewable edible, and does not compete with food, and has free
resources such as vegetable oils, cellulose, lignin, secondary hydroxyl groups. Approximately 90% of
starch, etc. [1]. The use of renewable raw materials fatty acids in castor oil are ricinoleic acid (C18:1),
can significantly contribute to sustainable develop- which have a hydroxyl functional group at the 12th
ment due to degradability and low toxicity of the re- carbon. This provides a hydroxyl value of between
sulting products. Conventional polyurethanes (PU) 160 and 180 mg KOH g1 [4, 5]. However, this low
are usually synthesized by a polyaddition reaction be- hydroxyl value along with the presence of secondary
tween polyols, which are petrochemical in origin and hydroxyls results in low functionality and low reac-
polyisocyanate, which forms urethane linkages [2] re- tivity [6, 7], leading to low crosslinking density, which
sulting in a crosslinked polymer. Nevertheless, poly- consequently produces semi-flexible and semi-rigid
urethanes can be obtained by using renewable sources materials among other limitations [8]. Sharma et al.
such as vegetable oils and can replace fossil fuel-de- [7] investigated flexible polyurethane foams synthe-
rived oligomers partially or totally. sized partially and completely from castor oil. They
showed that foams made from 100% of castor oil

*Corresponding
author, e-mail: demetrio.santos@ufabc.edu.br
BME-PT

927
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

were unstable and collapsed indicating the inferior et al. [19] characterized polyurethanic materials
reactivity of castor oil with isocyanate. The modifi- based on sulfonated lignin (unmodified and modified
cation of castor oil increases its hydroxyl value and by oxypropylation) and castor oil. The DMA results
hard segment composition to improve the rigidity, pointed out that the glass transition temperature (Tg)
physical and mechanical properties and crosslinking of the samples increased and thus the degree of cross-
density of the final PU products [9]. Nevertheless, the linking with the increase of hydroxyl groups derived
hydroxyl value of modified castor oil is limited. from different combinations of sulfonated lignin/
A renewable and promising source for sustainable sulfonated ligin oxypropylated/castor oil as polyols.
chemicals and bio-based polymeric materials is lignin. Cinelli et al. [20] characterized flexible polyure-
Its phenylpropanoic structure and high content of di- thanes foams from liquefied lignin and two different
verse functional groups (such as phenolic and aliphat- chain extenders: castor oil and poly(propylene gly-
ic hydroxyls, carbonyls, carboxyls) allow it to be col) (PPG). The single use of unmodified or modified
used as an alternative for polymer development es- castor oil as a polyol is already consolidated [3, 21].
pecially in the substitution of petroleum-based poly- Modified lignins have also been studied for this ap-
ols in polyurethane synthesis. Many tons of lignin are plication [22, 23]. However, the combination of mod-
generated as by-products of industrial processes such ified castor oil (MCO) and unmodified industrial
as pulp and paper. Most of the lignin extracted from lignin shows an interesting opportunity for renewable
pulp and paper operations is burned during pulp-spent and low cost polyols for preparation of PU.
liquor treatment. This offers energy recovery and re- The aim of this work is to develop and to character-
generation of pulping chemicals with less than 2% ize polyurethane obtained from renewable sources by
recovered for utilization as a chemical product [10]. using polyols including modified castor oil and un-
However, the amount of lignin produced exceeds the modified paper and pulp residue lignin. The influ-
requirements for energy generation. The type of pulp- ences of the basic chemistry reactions formed from
ing process determines the type of lignin industrially different combinations were investigated. Fourier
available because it unavoidably modifies the lignin transform infrared spectroscopy (FTIR) was used to
structure from that in the original feedstock. To in- identify functional groups of the polymers. The ther-
crease the potential applications of lignin in poly- mal and mechanical properties were studied using
meric materials, some chemical modifications have thermogravimetric analysis (TGA), dynamic mechan-
been developed [11, 12], but these add stages to the ical analysis (DMA) and tensile property measure-
process and/or raise their costs considerably. There- ments. Our research efforts focused on the develop-
fore, the direct use of industrial lignin is the most fa- ment of novel 100% renewable polyols able to syn-
vorable option because it is a relatively cheap raw thetize polyurethanes with a wide range of glass tran-
material. Unmodified lignin has poor stability [13] sition temperature (Tg) and mechanical properties.
and difficult melt processing [14], which make its di- The wide range of lignin-containing polyurethane
rect use uncompetitive. However, many studies have mechanical properties can make it suitable for the re-
focused on the incorporation of lignin in polymer placement of petroleum-based PU on several appli-
materials by blending it with synthetic or other bio- cations.
based polymers [1517].
Current studies have shown increasing interest in di- 2. Experimental section
versifying the sources of the hydroxyl groups. Mo- 2.1. Materials
hamed et al. [18] successfully synthetized an eco- Technical Kraft lignin (TKL) was obtained as a
friendly waterborne polyurethane dispersion, from byproduct of pulp and paper production. It was kind-
castor oil and aromatic polyamide sulfone, via copoly- ly supplied by Suzano Papel e Celulose (Suzano, SP,
merization reaction. Alternatively, combining lignin Brazil) with the following characteristics: brown
and castor oil as polyols for polyurethanic materials color, Mw = 3388 gmol1 (obtained by gel perme-
is promising. This can produce diversified materials ation chromatography) pH 8.1, solid content = 92.5%,
with varied properties and applications. de Oliveira ashes = 10% and total hydroxyl index equivalent to

928
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

Figure 1. Schematic drawing of castor oil modification reaction

243 mg KOH g1 (determined by 31P nuclear mag- beaker. Polyurethanes (PU-CO, PU-MCO1 and PU-
netic resonance). The TKL was dried at 80 C for 6 h MCO2) were prepared by mechanically mixing MDI
before polyol preparation. and the natural renewable source lignin-containing
Castor oil (CO) and two types of modified castor oil polyols (NCO/OH equivalent molar ratio of 1.2) for
(MCO1 and MCO2) were provided by CPA Brazil 2 min at 20 rpm in a 200 mL beaker [19]. Polyol com-
(Diadema, SP, Brazil) with a hydroxyl index of 159, positions are presented in Table 1.
237 and 286 mg KOH g1 and acid index of 0.95, 3.87 The mixed polyurethane was poured into the cavities
and 4.25 mg KOH g1, respectively (determined by of an open silicon mold, with cavity dimensions ac-
titration). MCO1 and MCO2 were synthetized using cording to ASTM D638-10 specimens Type I. The
a stainless steel industrial reaction kettle equipped cure was carried out at room temperature for 7 days.
with mechanical stirring, temperature monitoring, Figure 2 shows representative reaction schemes be-
cooling control system and N2 inlet. Castor oil was tween TKL/MDI/modified castor oil, which eluci-
modified by reacting with glycerol and Ca(OH)2 dates the urethane group formation and consequent
(1% of reactant total mass) at 230 C for 8 h. Varying polymeric structure.
castor oil/glycerol ratios gave modified castor oils
with different hydroxyl index. The reaction scheme 2.3. Characterizations
of castor oil modification is shown in Figure 1. 2.3.1. Fourier transform infrared spectroscopy
Diphenylmethane diisocyanate (MDI) was acquired (FTIR)
for polyurethane preparation from Kalium Chemical The spectroscopic measurements in the infrared re-
(So Paulo, SP, Brazil). It contained 30 and 32% gions (FTIR) were performed in a Thermo Nicolet
(minimum and maximum) NCO group values. Nexus 4700 spectrometer in transmittance mode; 10
scans were performed from 4000500 cm1 with a
2.2. Polyurethane preparation resolution of 4 cm1 in each sample.
Natural renewable source lignin-containing polyols
were obtained by adding 10, 20 or 30 wt% of tech- 2.3.2 Thermogravimetric analysis (TG)
nical Kraft lignin (TKL) with castor oil (CO) or with Thermogravimetric analysis (TG) was carried out
modified castor oils (MCO1 or MCO2). Lignin-con- using Netzsch equipment model STA 449F3. The
taining polyols were obtained by stirring 50 g of CO, samples (11.0 mg) were heated from 25 to 800 C
MCO1 or MCO2 and the respective lignin weight under nitrogen atmosphere and 50 mLmin1 flow
ratios. Polyol preparation was carried out under air at and a heating rate of 10 Cmin1.
room temperature, for 10 min at 80 rpm in a 200 mL
Table 1. Nomenclature and compositions of developed polyols according to TKL and oil wt%
PU-CO PU-CO/L10 PU-CO/L20 PU-CO/L30
TKL [wt%] 0 10 20 30
CO [wt%] 100 90 80 70
PU-MCO1 PU-MCO1/L10 PU-MCO1/L20 PU-MCO1/L30
TKL [wt%] 0 10 20 30
MCO1 [wt%] 100 90 80 70
PU-MCO2 PU-MCO2/L10 PU-MCO2/L20 PU-MCO2/L30
TKL [wt%] 0 10 20 30
MCO2 [wt%] 100 90 80 70

929
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

Figure 2. Schematic drawing of polyurethane groups formation from TKL/MDI/MCO reaction

2.3.3. Dynamic mechanical analysis (DMA) Type I test specimens using an Instron universal test-
Glass transition temperature (Tg) of renewable poly- ing machine model 5569 (CECS, Federal University
ol based polyurethanes was investigated by DMA of ABC, Santo Andr, SP, Brazil) with a crosshead
using a thermal analyzer model Pyris Diamond, speed of 50 mmmin1 and a non-contact extensome-
Perkin Elmer. Measurements were carried out in flex- ter (Instron SVE). Type I test specimens presented
ural mode (single cantilever) with a temperature range following dimensions: overall length (165 mm), width
from 50 to 150 C, frequency of 1 Hz, oscillation of narrow section (13 mm), overall thickness (5 mm)
amplitude of 10 mm and heating rate of 2 Cmin1. and gage length (50 mm).
The dimension of specimens was 30 mm12 mm
2 mm, which were obtained by casting PU into the 2.3.5. Scanning electron microscopy (SEM)
cavities of an open silicon mold. Storage modulus at After mechanical testing, SEM was used to observe
rubbery plateau (ER) was used to calculate cross- the cross sectional morphology of the fractured sur-
linking density. faces of the samples using a Jeol 6460LV scanning
electron microscope with an electron beam at an ac-
2.3.4. Mechanical properties celerating voltage of 25 kV. The samples were set on
The mechanical properties (5 specimens for each con- the SEM sample holder and sputter coated with a
dition) were tested according to ASTM D638 with thin layer of gold.

930
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

3. Results and discussion MCO2. The absence of NCO stretching at 2260 cm1
3.1. Fourier transform infrared spectroscopy indicates that the isocyanate groups have been entirely
(FTIR) reacted [19, 24, 27]. Other main absorption bands
Figure 3 shows the FTIR analysis for CO, MCO1 confirm polyurethane formation including absorp-
and MCO2 oils. All spectra present the characteristic tions in the region between 17041709 cm1 (car-
peak at 3400 cm1. This corresponds to the hydroxyl bonyl group hydrogen bonding to the urethane group)
group. MCO1 and MCO2 show increase in this band and 1215 cm1 (urethane linkages) [19, 23, 27].
intensity (Figure 3b), which is related to hydroxyl Figure 5 shows the FTIR spectra of lignin-containing
value increasing after modification. Other character- polyurethanes (varying the TKL wt% into the poly-
istic absorption bands are observed at 2920 cm1 urethanes) and of TKL (Figure 5b, 5d and 5f). The
(methyl), 2850 cm1 (methylene), 1740 cm1 (esters) band at 3425 cm1, TKL spectrum (Figure 5b, 5d and
and 1161 cm1 (COC) [23-26]. 5f), is characteristic of its aromatic and aliphatic OH
Figure 4 compares the FTIR spectra of PU-CO, PU- bond stretching [15, 28, 29]. The presence of lignin
MCO1 and PU-MCO2. A characteristic band of ure- in polyurethane resulted into the formation of a wider
thane stretching at 3330 cm1 is present on all sam- band in the 33303425 cm1 region due to merging
ples, which corresponds to hydrogen bonded NH of the bonded NH band (3330 cm1) with lignin OH
group in disordered form [19, 24]. The increasing in- bond stretching (3425 cm1) [27]. It can be observed
tensity of this band (Figure 4b) shows the increased in parts b, d and f of Figure 5 that the intensity of the
amount of urethane groups for PU-MCO1 and PU- merged region increases with the increasing of lignin
MCO2, which might be associated with reaction of quantity for all samples.
NCO with increased hydroxyl level of MCO1 and

Figure 3. FTIR spectra of vegetable oils (CO, MCO1 and MCO2) (a) from 3700 to 600 cm1 and (b) from 3700 to 3000 cm1

Figure 4. FTIR spectra of PU-CO, PU-MCO1 and PU-MCO2 (a) from 3700 to 600 cm1 and (b) from 3700 to 3000 cm1

931
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

Figure 5. FTIR spectra of PU-CO/lignin compositions (a) from 3700 to 600 cm1, (b) PU-CO/lignin and lignin (TKL) from
3700 to 3000 cm1; of PU-MCO1/lignin compositions, (c) from 3700 to 600 cm1, (d) PU-CO1/lignin and lignin
(TKL) from 3700 to 3000 cm1; and PU-MCO2/lignin (e) from 3700 to 600 cm1, (f) PU-CO2/lignin and lignin
(TKL) from 3700 to 3000 cm1

As can be seen in Figures 5a, 5c and 5e the spectra MCO2/lignin) [27]. The absence of NCO stretching
of lignin-containing polyurethanes show that as the at 2260 cm1 in all lignin-containing polyurethane
lignin content increases, the peak of the carbonyl spectra indicates that the isocyanate groups have
stretching vibration (around 1700 cm1) gradually been entirely reacted [19, 24, 27].
shifted to higher wavenumbers. Carbonyl bands of
polyurethane can be divided on three main regions: 3.2. Thermogravimetric analysis (TGA)
the hydrogen bonded carbonyl in ordered crystalline The TGA and derivative TG curves for castor oil-
domains at 17001709 cm1 [19, 24, 27], the hydro- based PU and castor oil/lignin-based PU are present-
gen bonded carbonyl in disordered amorphous con- ed in Figures 6 and 7. The stability of the PU is related
formations at 17141720 cm1 [19, 30] and the free to the hard segment nature (rigid aromatic ring of
carbonyl groups (non bonded) at 17311745 cm1 MDI and TKL), soft segment (introduced by flexible
[23, 31]. The carbonyl band shifting indicates that the chains of castor oil) and the molar ratio of the hard
presence of lignin induces the change from a hydro- segment to soft segment [32]. In general, the thermal
gen bonded ordered crystalline domain (PU-CO, PU- degradation of polyurethane occurs in a two to three
MCO1 and PU-MCO2) to a disordered amorphous steps, and the composition of the decomposed prod-
conformation and to the formation of free carbonyl ucts depends on the structure of the PU material [33].
groups (PU-CO/lignin, PU-MCO1/lignin and PU-

932
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

Figure 6. (a) TGA curves of PU-CO, PU-MCO1 and PU-MCO2 and (b) DTG curves of PU-CO, PU-MCO1 and PU-MCO2

Figure 7. (a) TGA curves of PU-CO and PU-CO/lignin, (b) DTG curves of PU-CO and PU-CO/lignin, (c) TGA curves of
PU-MCO1 and PU-MCO1/lignin, (d) DTG curves of PU-MCO1 and PU-MCO1/lignin, (e) TGA curves of PU-
MCO2 and PU-MCO2/lignin and (f) DTG curves of PU-MCO2 and PU-MCO2/lignin compositions

933
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

A first step decomposition is observed around 270 C Table 2. Thermogravimetric parameters of PU-CO/lignin,
and is related to the thermal decomposition of unsta- PU-MCO1/lignin and PU-MCO2/lignin
ble urethane bonds [27]. The second step occurs Tonset T50% Toffset Char residue
[C] [C] [C] [%]
above 320 C. The mass loss occurs at a fast rate in PU-CO 290.7 421.5 484.6 11.2
this step. It is associated to the soft segment and de- PU-CO/L10 287.7 434.8 485.9 17.0
pends on its structure and three-dimensional arrange- PU-CO/L20 284.2 438.1 480.9 22.2
ment [33]. The third one is above 380 C and is also PU-CO/L30 282.3 437.2 478.8 24.6
related to other remaining structures formed after the PU-MCO1 267.1 453.7 486.3 13.3
PU-MCO1/L10 265.6 452.6 487.9 19.5
second decomposition, diisocyanate and lignin aro-
PU-MCO1/L20 264.8 440.9 486.3 26.1
matic rings [1, 25, 34]. PU-MCO1/L30 264.3 444.9 477.8 26.5
Figure 6 shows the TGA and derivative curves of PU-MCO2 267.3 441.8 492.6 13.9
PU-CO, PU-MCO1 and PU-MCO2, respectively. The PU-MCO2/L10 263.6 440.9 488.7 18.4
PU-MCO2 and PU-MCO1 showed lower degrada- PU-MCO2/L20 258.0 429.0 480.3 20.3
PU-MCO2/L30 259.9 441.5 485.5 24.6
tion temperature for first step decomposition (max-
imum peaks at 289.3 and 294.2 C, respectively) and
presented higher mass loss compared to PU-CO mogravimetric parameters, including initial decom-
(320.0C) due to increasing urethane groups that were position temperature for degradation step (Tonset),
improved by castor oil modification. The second step final decomposition temperature for degradation step
decomposition may correspond to the chain scission (Toffset), temperature for 50% mass loss (T50%) and
of CO (PU-CO 393.6C, PU-MCO1 372.3 C, and % mass of remaining char at 750 C.
PU-MCO2 370.9 C), the soft polyurethane segment,
also observed by Hablot et al. [35]. In the third step 3.3. Dynamic mechanical analysis (DMA)
of decomposition, DTG peaks are shown at 448.5 C The evolution of tan as function of temperature is
for PU-CO, 457.1 C for PU-MCO1, and 459.4 C presented in Figure 8 and might be attributed to the
for PU-MCO2. glass transition temperature (Tg) [23]. A wide Tg range
Figure 7 shows TGA and DTG curves of PU-CO/ might be obtained for polyurethanes, which is de-
lignin, PU-MCO1/lignin and PU-MCO2/lignin com- pendent on polymer segment, hard and soft nature,
positions. The presence of lignin for all conditions and composition [19].
decreased the onset of thermal decomposition. The Figure 8a shows the glass transition temperature of
degradation onset in the 250290 C temperature PU-CO and PU-CO/lignin conditions. Tan values
range corresponds to the decomposition of unstable shifted to higher temperature as TKL wt% increases.
urethane bond from hard segment, but also the cleav- PU-CO presented the Tg at 0.76 C and is mainly
age of unstable ether linkages of lignin (Figure 7a, related to soft segments [19]. Increasing lignin con-
7c and 7e) [23]. Derivative curves for PU-CO/L20, tent shifted the Tg to a maximum value of 47.5 C for
PU-CO/L30 and PU-MCO1/L30 revealed two de- PU-CO/L30, mainly due to lignin rigid segment
composition peaks, as can be seen in Figure 7b and movements [19]. Lignin mass percentage and Tg shift-
7d. The two first stages, hard and soft segment degra- ing presented a linear dependence. Lignin mass frac-
dations, became close resulting in one larger peak, tion increasing also induced the tan peak broaden-
also observed by Zhang et al. [23]. The larger peaks ing, which implies in the sample heterogeneity and
have shifted to lower temperature and smaller inten- might be related to augment of molecular weight of
sity, compared to the second step decomposition of polyol component [36]. In this study, it might have
lignin free polyurethane, at 300400 C range. Lignin been induced by high molecular weight and hetero-
incorporation did not significantly affect the degra- geneous molecular structure of TKL added.
dation behavior but it rather increased the amount of Tan curves of PU-MCO1 and PU-MCO2 based
char formation, observed in TGA curves above polyurethanes are presented in Figures 8b and 8c. As
500 C. It was found that the char residue of lignin- general tendency, the Tg shifted to higher tempera-
containing polyurethanes increases as the lignin con- tures as TKL content increased. It can be seen in Fig-
centration increases. Table 2 presents additional ther- ures 8a, 8b and 8c that all lignin-containing polyure-

934
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

Figure 8. DMA curves (tan ) of (a) PU-CO/lignin, (b) PU-MCO1/lignin and (c) PU-MCO2/lignin

thanes show a single tan peak, which is an indica- to normal stress (tensile and flexural operation mode),
tive of polyol components miscibility. PU-CO is the Equation (1) [37]:
only condition that presented a small secondary peak,
ERl
which might be associated to a slight phase separa- o = 3RT (1)
tion between soft and hard segments in this compo-
sition. where R is the gas constant, ER is the storage mod-
It is of interest to evaluate the effects of CO modifi- ulus at rubbery plateau and T the absolute tempera-
cation and TKL addition on polyurethane crosslink- ture. Storage modulus at rubbery plateau and calcu-
ing density (), which can be calculated using stor- lated crosslinking density are presented in Table 3.
age modulus at rubbery plateau [36]. The crosslink- Castor oil modification increased polyurethane cross-
ing density can be obtained, for samples submitted linking density from 293 molm3 (PU-CO) to

Table 3. Storage modulus at rubbery plateau and crosslinking density of PU-CO/lignin, PU-MCO1/lignin and PU-
MCO2/lignin
PU-CO PU-CO/L10 PU-CO/L20 PU-CO/L30
ER [MPa] 2.5 6.1 6.9 13.2
[ molm3] 293 670 731 1360
PU-MCO1 PU-MCO1/L10 PU-MCO1/L20 PU-MCO1/L30
ER [MPa] 3.2 13.4 19.1 22.4
[ molm3] 331 1365 1919 2209
PU-MCO2 PU-MCO2/L10 PU-MCO2/L20 PU-MCO2/L30
ER [MPa] 5.4 9.1 13.1 14.6
[ molm3] 560 918 1288 1407

935
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

560 molm3 (PU-MCO2). The crosslinking density 3.4. Scanning electron microscopy (SEM)
increasing of PU-CO/lignin, PU-MCO1/lignin and Lignin-containing PU-COs presented a smooth tran-
PU-MCO2/lignin is a direct response of TKL wt% sition from a ductile fracture (smooth grooves, white
rising. However, PU-MCO1/lignin presented higher arrows) to a brittle one (sharp surfaces, black arrows)
crosslinking density than PU-MCO2/lignin, besides [38] near the TKL concentration value of 30 wt%,
the higher crosslinking density value for PU-MCO2 as illustrate in Figure 9a and 9b; for the PU-MCO1,
in comparison to PU-MCO1. there is a change in fracture mechanism at 20 wt%

Figure 9. SEM images of the fractured surface of the lignopolyurethanes with magnification of 10 000 (a, b and c) and
5000 (d, e and f)

936
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

TKL content (Figure 9c and 9d). This behavior is sim- mosets starts after chain scission and is not governed
ilar for PU-MCO2, where the value is 10 wt% lignin by the crosslinking density [40, 41], a comparison
content (Figure 9e and 9f). It reveals that for PU- of the increasing hydroxyl index and the ultimate
MCO1 and PU-MCO2 this behavior is anticipated strain is not feasible.
due to lower lignin content. The introduction of TKL in the polyol strongly af-
fected the mechanical behavior and increases the ul-
3.5. Mechanical properties timate tensile stress (Figure 10a) and Youngs mod-
Figure 10 summarizes the mechanical behavior of ulus (Figure 10c) of lignin-containing polyurethanes.
the polyurethanes. First, the increased hydroxyl index The maximum tensile stress was 23.50 MPa, and the
from CO to MCO1 and MCO2 substantially in- maximum Youngs modulus was 2.00 GPa for PU-
creased the ultimate tensile stress (Figure 10a) and MCO2/L30. The hardening effect caused by intro-
the Young modulus (Figure 10c) of PU-MCO1 and duction of TKL might be elucidated by two mecha-
PU-MCO2 versus PU-CO. Polyurethane mechanical nisms: i) Besides its participation as a co-monomer,
behavior is mainly determined by the crosslinking lignin has a reinforcement role derived from its char-
density given by the functionality of its reactants and acteristics (aromatic rigid structure and Youngs mod-
the stoichiometry between them [11, 39]. ulus between 2.314.65 GPa) [20, 39, 42, 43];
Figures 10b and 10d allow analyzing the effects of in- ii) Crosslinking density increasing [11, 20, 44].
creasing hydroxyl index on strain. Higher hydroxyl The ultimate strain for PU-CO and PU-MCO1 in-
indices decreased elastic strain (Figure 10d), which creased by adding TKL up to 20 and 10 wt%, respec-
might be associated to higher amounts of rigid seg- tively. As TKL weight contents rises for these com-
ments and higher crosslinking density in the polymer positions, the ultimate strain decreases. This ultimate
structure [11, 24]. Comparison of Figure 10b and strain trend inversion converges to ductile-to-brittle
10d highlights the presence of plastic deformation transition, as shown in Figure 9, and is related to more
in PU-CO, PU-MCO1 and PU-MCO2 before rup- rigid segments and consequently to glass transition
ture. Considering that plastic deformation of ther- temperature shifting. The ductile-to-brittle transition

Figure 10. Mechanical results of samples with different amounts of lignin: (a) ultimate stress, (b) ultimate strain, (c) Young
modulus and (d) elastic strain

937
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

occurred around 45 C for lignin-containing poly- [2] Huber G. W., Iborra S., Corma A.: Synthesis of trans-
urethanes, based on CO or MCOs, as can be observed portation fuels from biomass: Chemistry, catalysts, and
engineering. Chemical Reviews, 106, 40444098
in DMA results. (2006).
Results pointed out that the mechanical properties of DOI: 10.1021/cr068360d
lignin-containing polyurethane are in agreement with [3] Zhang L., Zhang M., Hu L., Zhou Y.: Synthesis of rigid
the mechanical properties of petroleum based poly- polyurethane foams with castor oil-based flame retar-
urethanes, such as polyurethane adhesives (Young dant polyols. Industrial Crops and Products, 52, 380
388 (2014).
modulus: 0.73 to 2.25 MPa; ultimate tensile stress: DOI: 10.1016/j.indcrop.2013.10.043
17.84 to 40.03 MPa) [45, 46] and crosslinked poly- [4] Karak N., Rana S., Cho J. W.: Synthesis and character-
urethanes (typical ultimate tensile stress: 1.70 to ization of castor-oil-modified hyperbranched polyure-
41.00 MPa) [4749]. PU-MCO2/L30 presented the thanes. Journal of Applied Polymer Science, 112, 736
highest Young modulus in comparison with recent 743 (2009).
DOI: 10.1002/app.29468
published results concerning 100% bio-based poly- [5] Zhang M., Pan H., Zhang L., Hu L., Zhou Y.: Study of
ols based polyurethanes [23, 50, 51]. the mechanical, thermal properties and flame retardan-
cy of rigid polyurethane foams prepared from modified
4. Conclusions castor-oil-based polyols. Industrial Crops and Products,
Polyurethanes based on renewable raw materials 59, 135143 (2014).
DOI: 10.1016/j.indcrop.2014.05.016
(technical Kraft lignin, castor oil and modified castor [6] Gallezot P.: Conversion of biomass to selected chemical
oil) were prepared. The results indicated that the start- products. Chemical Society Reviews, 41, 15381558
ing materials change the polymers properties. The (2012).
oil modification process improved hydroxyl concen- DOI: 10.1039/C1CS15147A
tration and lead to polyurethanes with enhanced me- [7] Sharma C., Kumar S., Unni A. R., Aswal V. K., Rath S.
K., Harikrishnan G.: Foam stability and polymer phase
chanical properties versus ones synthesized from un- morphology of flexible polyurethane foams synthesized
modified castor oil. The introduction of TKL in- from castor oil. Journal of Applied Polymer Science,
creased the glass transition temperature of the mate- 131, 84208427 (2014).
rials from 0.76 C for PU-CO to 47.5 C for PU- DOI: 10.1002/app.40668
CO/L30 and had a twofold effect on the mechanical [8] Mutlu H., Meier M. A. R.: Castor oil as a renewable re-
source for the chemical industry. European Journal of
properties. The reinforcement co-monomer and fur- Lipid Science and Technology, 112, 1030 (2010).
ther increase in hydroxyl content led to a higher DOI: 10.1002/ejlt.200900138
crosslinking density. The products show the feasibil- [9] Mosiewicki M. A., DellArciprete G. A., Aranguren M.
ity of developing polyurethane-type materials with I., Marcovich N. E.: Polyurethane foams obtained from
large property range, by using an industrial low cost, castor oil-based polyol and filled with wood flour. Jour-
nal of Composite Materials, 43, 30573072 (2009).
unmodified and largely available residue combined DOI: 10.1177/0021998309345342
with no edible renewable source oil. [10] Lora J.: Industrial commercial lignins: Sources, prop-
erties and applications. in Monomers, polymers and
Acknowledgements composites from renewable resources (eds: Belgacem
The authors thank FAPESP/CNPq and acknowledge UFABC M. N., Gandini A.) Elsevier, Amsterdam, 225241
(Federal University of ABC) for the doctoral fellowship of (2008).
L.T. and A.N., the master fellowship of C.B. and G.S. and DOI: 10.1016/B978-0-08-045316-3.00010-7
nancial support. [11] Duval A., Lawoko M.: A review on lignin-based poly-
meric, micro- and nano-structured materials. Reactive
and Functional Polymers, 85, 7896 (2014).
References DOI: 10.1016/j.reactfunctpolym.2014.09.017
[1] Bernardini J., Cinelli P., Anguillesi I., Coltelli M-B., [12] Kandula M., Schwenke T., Friebel S., Salthammer T.:
Lazzeri A.: Flexible polyurethane foams green produc- Effect of ball milling on lignin polyesterification with
tion employing lignin or oxypropylated lignin. Euro- -caprolactone. Holzforschung, 69, 297302 (2015).
pean Polymer Journal, 64, 147156 (2015). DOI: 10.1515/hf-2014-0053
DOI: 10.1016/j.eurpolymj.2014.11.039 [13] Braun J. L., Holtman K. M., Kadla J. F.: Lignin-based
carbon fibers: Oxidative thermostabilization of kraft
lignin. Carbon, 43, 385394 (2005).
DOI: 10.1016/j.carbon.2004.09.027

938
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

[14] Norberg I., Nordstrm Y., Drougge R., Gellerstedt G., [26] Wong C. S., Badri K. H.: Chemical analyses of palm
Sjholm E.: A new method for stabilizing softwood kernel oil-based polyurethane prepolymer. Materials
kraft lignin fibers for carbon fiber production. Journal Science and Applications, 3, 7886 (2012).
of Applied Polymer Science, 128, 38243830 (2013). DOI: 10.4236/msa.2012.32012
DOI: 10.1002/app.38588 [27] Luo X., Mohanty A., Misra M.: Lignin as a reactive re-
[15] Gordobil O., Delucis R., Egs I., Labidi J.: Kraft lignin inforcing filler for water-blown rigid biofoam compos-
as filler in PLA to improve ductility and thermal prop- ites from soy oil-based polyurethane. Industrial Crops
erties. Industrial Crops and Products, 72, 4653 (2015). and Products, 47, 1319 (2013).
DOI: 10.1016/j.indcrop.2015.01.055 DOI: 10.1016/j.indcrop.2013.01.040
[16] Schorr D., Diouf P. N., Stevanovic T.: Evaluation of in- [28] Garca A., Erdocia X., Gonzlez M. A., Labidi J.: Effect
dustrial lignins for biocomposites production. Industrial of ultrasound treatment on the physicochemical prop-
Crops and Products, 52, 6573 (2014). erties of alkaline lignin. Chemical Engineering and Pro-
DOI: 10.1016/j.indcrop.2013.10.014 cessing: Process Intensification, 62, 150158 (2012).
[17] Spiridon I., Leluk K., Resmerita A. M., Darie R. N.: DOI: 10.1016/j.cep.2012.07.011
Evaluation of PLAlignin bioplastics properties before [29] Nada A-A. M. A., Yousef M. A., Shaffei K. A., Salah
and after accelerated weathering. Composites Part B: A. M.: Infrared spectroscopy of some treated lignins.
Engineering, 69, 342349 (2015). Polymer Degradation and Stability, 62, 157163 (1998).
DOI: 10.1016/j.compositesb.2014.10.006 DOI: 10.1016/S0141-3910(97)00273-5
[18] Mohamed H. A., Badran B. M., Rabie A. M., Morsi S. [30] Pietrzak K., Kirpluks M., Cabulis U., Ryszkowska J.:
M. M.: Synthesis and characterization of aqueous (poly- Effect of the addition of tall oil-based polyols on the
urethane/aromatic polyamide sulfone) copolymer dis- thermal and mechanical properties of ureaurethane elas-
persions from castor oil. Progress in Organic Coatings, tomers. Polymer Degradation and Stability, 108, 201
77, 965974 (2014). 211 (2014).
DOI: 10.1016/j.porgcoat.2014.01.026 DOI: 10.1016/j.polymdegradstab.2014.03.038
[19] de Oliveira F. D., Ramires E. C., Frollini E., Belgacem [31] Sormana J-L., Meredith J. C.: High-throughput discov-
M. N.: Lignopolyurethanic materials based on oxy- ery of structuremechanical property relationships for
propylated sodium lignosulfonate and castor oil blends. segmented poly(urethaneurea)s. Macromolecules, 37,
Industrial Crops and Products, 72, 110 (2015). 21862195 (2004).
DOI: 10.1016/j.indcrop.2015.01.023 DOI: 10.1021/ma035385v
[20] Cinelli P., Anguillesi I., Lazzeri A.: Green synthesis of [32] Corcuera M. A., Rueda L., Fernandez DArlas B., Ar-
flexible polyurethane foams from liquefied lignin. Eu- belaiz A., Marieta C., Mondragon I., Eceiza A.: Mi-
ropean Polymer Journal, 49, 11741184 (2013). crostructure and properties of polyurethanes derived
DOI: 10.1016/j.eurpolymj.2013.04.005 from castor oil. Polymer Degradation and Stability, 95,
[21] Yeganeh H., Mehdizadeh M. R.: Synthesis and proper- 21752184 (2010).
ties of isocyanate curable millable polyurethane elas- DOI: 10.1016/j.polymdegradstab.2010.03.001
tomers based on castor oil as a renewable resource poly- [33] Chattopadhyay D. K., Webster D. C.: Thermal stability
ol. European Polymer Journal, 40, 12331238 (2004). and flame retardancy of polyurethanes. Progress in
DOI: 10.1016/j.eurpolymj.2003.12.013 Polymer Science, 34, 10681133 (2009).
[22] Huo S-P., Nie M-C., Kong Z-W., Wu G-M., Chen J.: DOI: 10.1016/j.progpolymsci.2009.06.002
Crosslinking kinetics of the formation of lignin-aminat- [34] Dsouza J., Camargo R., Yan N.: Polyurethane foams
ed polyol-based polyurethane foam. Journal of Applied made from liquefied bark-based polyols. Journal of Ap-
Polymer Science, 125, 152157 (2012). plied Polymer Science, 131, 110 (2014).
DOI: 10.1002/app.35401 DOI: 10.1002/app.40599
[23] Zhang C., Wu H., Kessler M. R.: High bio-content poly- [35] Hablot E., Zheng D., Bouquey M., Avrous L.: Poly-
urethane composites with urethane modified lignin as urethanes based on castor oil: Kinetics, chemical, me-
filler. Polymer, 69, 5257 (2015). chanical and thermal properties. Macromolecular Ma-
DOI: 10.1016/j.polymer.2015.05.046 terials and Engineering, 293, 922929 (2008).
[24] Das S., Pandey P., Mohanty S., Nayak S. K.: Influence DOI: 10.1002/mame.200800185
of NCO/OH and transesterified castor oil on the struc- [36] Kang S. M., Lee S. J., Kim B. K.: Shape memory
ture and properties of polyurethane: Synthesis and char- polyurethane foams. Express Polymer Letters, 6, 63
acterization. Materials Express, 5, 377389 (2015). 69 (2012).
DOI: 10.1166/mex.2015.1254 DOI: 10.3144/expresspolymlett.2012.7
[25] Gurunathan T., Mohanty S., Nayak S. K.: Isocyanate [37] Calvo-Correas T., Gabilondo N., Alonso-Varona A.,
terminated castor oil-based polyurethane prepolymer: Palomares T., Corcuera M. A., Eceiza A.: Shape-mem-
Synthesis and characterization. Progress in Organic ory properties of crosslinked biobased polyurethanes.
Coatings, 80, 3948 (2015). European Polymer Journal, 78, 253263 (2016).
DOI: 10.1016/j.porgcoat.2014.11.017 DOI: 10.1016/j.eurpolymj.2016.03.030

939
Tavares et al. eXPRESS Polymer Letters Vol.10, No.11 (2016) 927940

[38] Nair M. N. R., Sukumar P., Jayashree V., Nair M. R. G.: [45] Banea M. D., da Silva L. F. M., Carbas R. J. C., Campil-
Mechanical properties and fractography of block copoly- ho R. D. S. G.: Mechanical and thermal characterization
mers based on NR and MDI-based polyurethanes. Poly- of a structural polyurethane adhesive modified with
mer Bulletin, 65, 8396 (2010). thermally expandable particles. International Journal of
DOI: 10.1007/s00289-010-0251-8 Adhesion and Adhesives, 54, 191199 (2014).
[39] Ionescu M.: Chemistry and technology of polyols for DOI: 10.1016/j.ijadhadh.2014.06.008
polyurethane. Rapra, Shawbury (2007). [46] Clauss S., Gabriel J., Karbach A., Matner M., Niemz
[40] Jang B. Z., Pater R. H., Soucek M. D., Hinkley J. A.: P.: Influence of the adhesive formulation on the me-
Plastic deformation mechanisms in polyimide resins chanical properties and bonding performance of poly-
and their semi-interpenetrating networks. Journal of urethane prepolymers. Holzforschung, 65, 835844
Polymer Science Part B: Polymer Physics, 30, 643654 (2011).
(1992). DOI: 10.1515/HF.2011.095
DOI: 10.1002/polb.1992.090300701 [47] Prisacariu C.: Polyurethane elastomers: From morphol-
[41] Mullins M. J., Liu D., Sue H-J.: Mechanical properties ogy to mechanical aspects. Springer, Viena (2011).
of thermosets. in Thermosets: Structure, properties and DOI: 10.1007/978-3-7091-0514-6
applications (ed.: Guo Q.) Woodhead, Oxford (2012). [48] Wang H. H., Mou J., Ni Y. H., Fei G. Q., Si C. L., Zou
DOI: 10.1533/9780857097637.1.28 J.: Phase behavior, interaction and properties of acetic
[42] Elder T.: Quantum chemical determination of Youngs acid lignin-containing polyurethane films coupled with
modulus of lignin. Calculations on a O-4 model com- aminopropyltriethoxy silane. Express Polymer Letters,
pound. Biomacromolecules, 8, 36193627 (2007). 7, 443455 (2013).
DOI: 10.1021/bm700663y DOI: 10.3144/expresspolymlett.2013.41
[43] Li Y., Ragauskas A. J.: Kraft lignin-based rigid poly- [49] Kupka V., Vojtova L., Fohlerova Z., Jancar J.: Solvent
urethane foam. Journal of Wood Chemistry and Tech- free synthesis and structural evaluation of polyurethane
nology, 32, 210224 (2012). films based on poly(ethylene glycol) and poly(capro-
DOI: 10.1080/02773813.2011.652795 lactone). Express Polymer Letters, 10, 479492 (2016).
[44] Mahmood N., Yuan Z., Schmidt J., Xu C.: Preparation DOI: 10.3144/expresspolymlett.2016.46
of bio-based rigid polyurethane foam using hydrolyti- [50] Lee A., Deng Y.: Green polyurethane from lignin and
cally depolymerized Kraft lignin via direct replacement soybean oil through non-isocyanate reactions. Euro-
or oxypropylation. European Polymer Journal, 68, 19 pean Polymer Journal, 63, 6773 (2014).
(2015). DOI: 10.1016/j.eurpolymj.2014.11.023
DOI: 10.1016/j.eurpolymj.2015.04.030 [51] Madhukar B. S., Bhadre Gowda D. G., Annadurai V.,
Somashekar R., Siddaramaiah: Phase behaviors of
PU/SPI green composites using SAXS profiles. Ad-
vances in Polymer Technology, 35, 21526/121526/10
(2016).
DOI: 10.1002/adv.21526

940

Potrebbero piacerti anche